Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

eIF4A1 enhances LARP1-mediated translational repression during mTORC1 inhibition

Abstract

Eukaryotic translation initiation factor (eIF)4A—a DEAD-box RNA-binding protein—plays an essential role in translation initiation. Recent reports have suggested helicase-dependent and helicase-independent functions for eIF4A, but the multifaceted roles of eIF4A have not been fully explored. Here we show that eIF4A1 enhances translational repression during the inhibition of mechanistic target of rapamycin complex 1 (mTORC1), an essential kinase complex controlling cell proliferation. RNA pulldown followed by sequencing revealed that eIF4A1 preferentially binds to mRNAs containing terminal oligopyrimidine (TOP) motifs, whose translation is rapidly repressed upon mTORC1 inhibition. This selective interaction depends on a La-related RNA-binding protein, LARP1. Ribosome profiling revealed that deletion of EIF4A1 attenuated the translational repression of TOP mRNAs upon mTORC1 inactivation. Moreover, eIF4A1 increases the interaction between TOP mRNAs and LARP1 and, thus, ensures stronger translational repression upon mTORC1 inhibition. Our data show the multimodality of eIF4A1 in modulating protein synthesis through an inhibitory binding partner and provide a unique example of the repressive role of a universal translational activator.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: eIF4A1 preferentially interacts with TOP mRNAs.
Fig. 2: LARP1 shows preferential binding to eIF4A1.
Fig. 3: Depletion of LARP1 eliminates the biased interaction between eIF4A1 and TOP mRNAs.
Fig. 4: EIF4A1-KO cells exhibit reduced translational repression of TOP mRNAs upon mTOR inhibition.
Fig. 5: eIF4A1 enhances the interaction between LARP1 and TOP RNAs.
Fig. 6: EIF4A1-KO cells are resistant to cytotoxicity mediated by mTOR inhibition.
Fig. 7: eIF4A2 is not required for the enhancement of LARP1-mediated translational repression.

Similar content being viewed by others

Data availability

The human genome hg38 and GENCODE release 32 annotations were obtained from UCSC Genome Browser (https://genome.ucsc.edu). The RNA pulldown-seq, ribosome profiling and RNA-seq (GSE184247) data used in this study were deposited in the National Center for Biotechnology Information. The MS (JPST002319) data used in this study were deposited in the Japan Proteome Standard Repository (jPOSTrepo). Source data are provided with this paper.

Code availability

Source codes of softwares for ribosome profiling data analysis were at GitHub (https://github.com/ingolia-lab/RiboSeq). The codes for data analysis of deep sequencing and MS have been deposited in Zenodo at https://doi.org/10.5281/zenodo.10644574 (ref. 69).

References

  1. Shah, P., Ding, Y., Niemczyk, M., Kudla, G. & Plotkin, J. B. Rate-limiting steps in yeast protein translation. Cell 153, 1589–1601 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Yan, X., Hoek, T. A., Vale, R. D. & Tanenbaum, M. E. Dynamics of translation of single mRNA molecules in vivo. Cell 165, 976–989 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Sonenberg, N. & Hinnebusch, A. G. Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136, 731–745 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Hinnebusch, A. G. The scanning mechanism of eukaryotic translation initiation. Annu. Rev. Biochem. 83, 779–812 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Brito Querido, J. et al. Structure of a human 48S translational initiation complex. Science 369, 1220–1227 (2020).

    Article  CAS  PubMed  Google Scholar 

  6. Rozen, F. et al. Bidirectional RNA helicase activity of eucaryotic translation initiation factors 4A and 4F. Mol. Cell. Biol. 10, 1134–1144 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Rogers, G. W., Richter, N. J. & Merrick, W. C. Biochemical and kinetic characterization of the RNA helicase activity of eukaryotic initiation factor 4A. J. Biol. Chem. 274, 12236–12244 (1999).

    Article  CAS  PubMed  Google Scholar 

  8. García-García, C., Frieda, K. L., Feoktistova, K., Fraser, C. S. & Block, S. M. RNA BIOCHEMISTRY. Factor-dependent processivity in human eIF4A DEAD-box helicase. Science 348, 1486–1488 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Andreou, A. Z. & Klostermeier, D. The DEAD-box helicase eIF4A: paradigm or the odd one out. RNA Biol. 10, 19–32 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Waldron, J. A. et al. mRNA structural elements immediately upstream of the start codon dictate dependence upon eIF4A helicase activity. Genome Biol. 20, 300 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Steinberger, J. et al. Identification and characterization of hippuristanol-resistant mutants reveals eIF4A1 dependencies within mRNA 5′ leader regions. Nucleic Acids Res. 48, 9521–9537 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Svitkin, Y. V. et al. The requirement for eukaryotic initiation factor 4A (eIF4A) in translation is in direct proportion to the degree of mRNA 5′ secondary structure. RNA 7, 382–394 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Sen, N. D., Zhou, F., Ingolia, N. T. & Hinnebusch, A. G. Genome-wide analysis of translational efficiency reveals distinct but overlapping functions of yeast DEAD-box RNA helicases Ded1 and eIF4A. Genome Res. 25, 1196–1205 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Iwasaki, S., Floor, S. N. & Ingolia, N. T. Rocaglates convert DEAD-box protein eIF4A into a sequence-selective translational repressor. Nature 534, 558–561 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Yourik, P. et al. Yeast eIF4A enhances recruitment of mRNAs regardless of their structural complexity. eLife 6, e31476 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Sokabe, M. & Fraser, C. S. A helicase-independent activity of eIF4A in promoting mRNA recruitment to the human ribosome. Proc. Natl Acad. Sci. USA 114, 6304–6309 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Tauber, D. et al. Modulation of RNA condensation by the DEAD-box protein eIF4A. Cell 180, 411–426.e16 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Farache, D., Antine, S. P. & Lee, A. S. Y. Moonlighting translation factors: multifunctionality drives diverse gene regulation. Trends Cell Biol. 32, 762–772 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Saxton, R. A. & Sabatini, D. M. mTOR signaling in growth, metabolism, and disease. Cell 169, 361–371 (2017).

    Article  CAS  PubMed  Google Scholar 

  20. Levy, S., Avni, D., Hariharan, N., Perry, R. P. & Meyuhas, O. Oligopyrimidine tract at the 5′ end of mammalian ribosomal protein mRNAs is required for their translational control. Proc. Natl Acad. Sci. USA 88, 3319–3323 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Jefferies, H. B., Reinhard, C., Kozma, S. C. & Thomas, G. Rapamycin selectively represses translation of the ‘polypyrimidine tract’ mRNA family. Proc. Natl Acad. Sci. USA 91, 4441–4445 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Hsieh, A. C. et al. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485, 55–61 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Thoreen, C. C. et al. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485, 109–113 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Meyuhas, O. & Kahan, T. The race to decipher the top secrets of TOP mRNAs. Biochim. Biophys. Acta 1849, 801–811 (2015).

    Article  CAS  PubMed  Google Scholar 

  25. Fonseca, B. D. et al. La-related Protein 1 (LARP1) represses terminal oligopyrimidine (TOP) mRNA translation downstream of mTOR complex 1 (mTORC1). J. Biol. Chem. 290, 15996–16020 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Lahr, R. M. et al. La-related protein 1 (LARP1) binds the mRNA cap, blocking eIF4F assembly on TOP mRNAs. eLife 6, e24146 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Philippe, L., Vasseur, J. J., Debart, F. & Thoreen, C. C. La-related protein 1 (LARP1) repression of TOP mRNA translation is mediated through its cap-binding domain and controlled by an adjacent regulatory region. Nucleic Acids Res. 46, 1457–1469 (2018).

    Article  CAS  PubMed  Google Scholar 

  28. Philippe, L., van den Elzen, A. M. G., Watson, M. J. & Thoreen, C. C. Global analysis of LARP1 translation targets reveals tunable and dynamic features of 5′ TOP motifs. Proc. Natl Acad. Sci. USA 117, 5319–5328 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Smith, E. M. et al. The mTOR regulated RNA-binding protein LARP1 requires PABPC1 for guided mRNA interaction. Nucleic Acids Res. 49, 458–478 (2020).

    Article  PubMed Central  Google Scholar 

  30. Lahr, R. M. et al. The La-related protein 1-specific domain repurposes HEAT-like repeats to directly bind a 5′ TOP sequence. Nucleic Acids Res. 43, 8077–8088 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jia, J. J. et al. mTORC1 promotes TOP mRNA translation through site-specific phosphorylation of LARP1. Nucleic Acids Res. 49, 3461–3489 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Linder, P. & Jankowsky, E. From unwinding to clamping—the DEAD box RNA helicase family. Nat. Rev. Mol. Cell Biol. 12, 505–516 (2011).

    Article  CAS  PubMed  Google Scholar 

  33. Emmott, E. & Goodfellow, I. Identification of protein interaction partners in mammalian cells using SILAC-immunoprecipitation quantitative proteomics. J. Vis. Exp. 6, 51656 (2014).

    Google Scholar 

  34. Iwasaki, S. et al. The translation inhibitor rocaglamide targets a bimolecular cavity between eIF4A and polypurine RNA. Mol. Cell 73, 738–748.e9 (2019).

    Article  CAS  PubMed  Google Scholar 

  35. Hellen, C. U. & Sarnow, P. Internal ribosome entry sites in eukaryotic mRNA molecules. Genes Dev. 15, 1593–1612 (2001).

    Article  CAS  PubMed  Google Scholar 

  36. Liu, J., Xu, Y., Stoleru, D. & Salic, A. Imaging protein synthesis in cells and tissues with an alkyne analog of puromycin. Proc. Natl Acad. Sci. USA 109, 413–418 (2012).

    Article  CAS  PubMed  Google Scholar 

  37. Bordeleau, M. E. et al. Functional characterization of IRESes by an inhibitor of the RNA helicase eIF4A. Nat. Chem. Biol. 2, 213–220 (2006).

    Article  CAS  PubMed  Google Scholar 

  38. Lindqvist, L. et al. Selective pharmacological targeting of a DEAD box RNA helicase. PLoS ONE 3, e1583 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Galicia-Vázquez, G., Cencic, R., Robert, F., Agenor, A. Q. & Pelletier, J. A cellular response linking eIF4AI activity to eIF4AII transcription. RNA 18, 1373–1384 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  40. Thoreen, C. C. et al. An ATP-competitive mammalian target of rapamycin inhibitor reveals rapamycin-resistant functions of mTORC1. J. Biol. Chem. 284, 8023–8032 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Feldman, M. E. et al. Active-site inhibitors of mTOR target rapamycin-resistant outputs of mTORC1 and mTORC2. PLoS Biol. 7, e38 (2009).

    Article  PubMed  Google Scholar 

  42. Aoki, K. et al. LARP1 specifically recognizes the 3′ terminus of poly(A) mRNA. FEBS Lett. 587, 2173–2178 (2013).

    Article  CAS  PubMed  Google Scholar 

  43. Gentilella, A. et al. Autogenous control of 5′TOP mRNA stability by 40S ribosomes. Mol. Cell 67, 55–70.e4 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Merrick, W. C. & Pavitt, G. D. Protein synthesis initiation in eukaryotic cells. Cold Spring Harb. Perspect. Biol. 10, a033092 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Wek, R. C. Role of eIF2α kinases in translational control and adaptation to cellular stress. Cold Spring Harb. Perspect. Biol. 10, a032870 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Kenner, L. R. et al. eIF2B-catalyzed nucleotide exchange and phosphoregulation by the integrated stress response. Science 364, 491–495 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Kashiwagi, K. et al. Structural basis for eIF2B inhibition in integrated stress response. Science 364, 495–499 (2019).

    Article  CAS  PubMed  Google Scholar 

  48. Duncan, R., Milburn, S. C. & Hershey, J. W. Regulated phosphorylation and low abundance of HeLa cell initiation factor eIF-4F suggest a role in translational control. Heat shock effects on eIF-4F. J. Biol. Chem. 262, 380–388 (1987).

    Article  CAS  PubMed  Google Scholar 

  49. Cassidy, K. C. et al. Capturing the mechanism underlying TOP mRNA binding to LARP1. Structure 27, 1771–1781.e5 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hua, H. et al. Targeting mTOR for cancer therapy. J. Hematol. Oncol. 12, 71 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Ducker, G. S. et al. Incomplete inhibition of phosphorylation of 4E-BP1 as a mechanism of primary resistance to ATP-competitive mTOR inhibitors. Oncogene 33, 1590–1600 (2014).

    Article  CAS  PubMed  Google Scholar 

  52. Jastrzebski, K. et al. Integrative modeling identifies key determinants of inhibitor sensitivity in breast cancer cell lines. Cancer Res. 78, 4396–4410 (2018).

    Article  CAS  PubMed  Google Scholar 

  53. Rodrik-Outmezguine, V. S. et al. Overcoming mTOR resistance mutations with a new-generation mTOR inhibitor. Nature 534, 272–276 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Alain, T. et al. eIF4E/4E-BP ratio predicts the efficacy of mTOR targeted therapies. Cancer Res. 72, 6468–6476 (2012).

    Article  CAS  PubMed  Google Scholar 

  55. Huang, D. W., Sherman, B. T. & Lempicki, R. A. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat. Protoc. 4, 44–57 (2009).

    Article  CAS  PubMed  Google Scholar 

  56. Bailey, T. L. & Elkan, C. Fitting a mixture model by expectation maximization to discover motifs in biopolymers. Proc. Int. Conf. Intell. Syst. Mol. Biol. 2, 28–36 (1994).

    CAS  PubMed  Google Scholar 

  57. Chen, M. et al. Dual targeting of DDX3 and eIF4A by the translation inhibitor rocaglamide A. Cell Chem. Biol. 28, 475–486.e8 (2021).

    Article  CAS  PubMed  Google Scholar 

  58. Rubio, C. A. et al. Transcriptome-wide characterization of the eIF4A signature highlights plasticity in translation regulation. Genome Biol. 15, 476 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  59. McGlincy, N. J. & Ingolia, N. T. Transcriptome-wide measurement of translation by ribosome profiling. Methods 126, 112–129 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Mito, M., Mishima, Y. & Iwasaki, S. Protocol for disome profiling to survey ribosome collision in humans and zebrafish. STAR Protoc. 1, 100168 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Chen, S., Zhou, Y., Chen, Y. & Gu, J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34, i884–i890 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  63. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Huang, D. W., Sherman, B. T. & Lempicki, R. A. Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res. 37, 1–13 (2009).

    Article  PubMed  Google Scholar 

  65. Gandin, V. et al. nanoCAGE reveals 5′ UTR features that define specific modes of translation of functionally related MTOR-sensitive mRNAs. Genome Res. 26, 636–648 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Wiśniewski, J. R., Zougman, A., Nagaraj, N. & Mann, M. Universal sample preparation method for proteome analysis. Nat. Methods 6, 359–362 (2009).

    Article  PubMed  Google Scholar 

  67. Saeki, N. et al. N-terminal deletion of Swi3 created by the deletion of a dubious ORF YJL175W mitigates protein burden effect in S. cerevisiae. Sci. Rep. 10, 9500 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Wu, Q. et al. Selective translation of epigenetic modifiers affects the temporal pattern and differentiation of neural stem cells. Nat. Commun. 13, 470 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Shichino, Y. & Iwasaki, S. Custom scripts for 'eIF4A1 enhances LARP1-mediated translational repression during mTORC1 inhibition'. Zenodo https://doi.org/10.5281/zenodo.10644574 (2024).

Download references

Acknowledgements

We are grateful to all the members of the Iwasaki laboratory for constructive discussions, technical help and critical reading of the manuscript. We also thank K. Dodo and M. Sodeoka for the Pharos FX imaging scanner, the Support Unit for Bio-Material Analysis, the RIKEN CBS Research Resources Division for MS and Sanger sequencing and the HOKUSAI SailingShip supercomputer facility at RIKEN for computational support. Hippuristanol was a kind gift from J. Tanaka. This work was supported by the Ministry of Education, Culture, Sports, Science and Technology (JP20H05784 to S.I., JP21H05734 and JP23H04268 to Y.S. and JP21H05281 to T.I.); the Japan Society for the Promotion of Science (JP17H04998 to S.I., JP19J00920 and JP21K15023 to Y.S., JP19H03172 to T.I., JP18K14644 to K. Kashiwagi and JP20H03426 and JP20K21566 to K. Kuba); the Japan Agency for Medical Research and Development (JP20gm1410001 to S.I. and T.I. and JP23gm6910005 to Y.S.) and RIKEN (Pioneering Projects ‘Biology of Intracellular Environments’ to Y.S., S.I. and T.I.; the BDR Structural Cell Biology Project, the Pioneering Projects ‘Dynamic Structural Biology’ to T.I.; Special Postdoctoral Researchers to Y.S.; and Incentive Research Projects to Y.S.). The DNA libraries were sequenced by the Vincent J. Coates Genomics Sequencing Laboratory at the Universtiy of California Berkeley, which is supported by a National Institutes for Health Instrumentation Grant (S10 OD018174). Y.S. was a recipient of a Japan Society for the Promotion of Science Research Fellow (PD) and was supported by the RIKEN Special Postdoctoral Researchers Program.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: Y.S. and S.I.; methodology: Y.S. and S.I.; formal analysis: Y.S. and T.Y.; investigation: Y.S., T.Y., M.M., K. Kashiwagi, M.T. and S.I.; resources: N.T.I.; writing—original draft: Y.S. and S.I.; writing—review and editing: Y.S., T.Y., M.M., K. Kashiwagi, M.T., T.I., N.T.I., K. Kuba and S.I.; visualization: Y.S.; supervision: T.I., N.T.I., K. Kuba and S.I.; funding acquisition: K. Kashiwagi, K. Kuba, T.I., Y.S. and S.I.

Corresponding authors

Correspondence to Yuichi Shichino or Shintaro Iwasaki.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Sara Osman and Carolina Perdigoto were the primary editors on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team. Peer reviewer reports are available.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 RNA pulldown-Seq with eIF4A1, Related to Fig. 1.

a, Western blotting of SBP-tagged eIF4A1 expressed in HEK293 cells. β-Actin was used as the loading control. b, CBB staining of purified SBP-eIF4A1 protein for RNA pulldown-Seq analysis.

Extended Data Fig. 2 Protein binding preference of eIF4A1, Related to Fig. 2.

a, Volcano plot of the heavy/light ratio. eIF4A1-bound proteins (defined as those with FDR < 0.05 and log2[heavy/light] > 1) are highlighted. b, Electropherogram of RNAs from the lysates used to obtain the data shown in Fig. 2c. c, Interactions between SBP-eIF4A1 and the series of N-terminal and C-terminal truncation mutants of LARP1. Naïve and SBP-EIF4A1 cells were transfected with 3 × FLAG-6 × His-tagged LARP1 and its variants. The domain structure of LARP1 is shown at the top.

Extended Data Fig. 3 Translational repression upon mTOR inhibition is attenuated in EIF4A1 KO cells, Related to Fig. 4.

a, Western blotting of eIF4A1 and eIF4A2 in the SBP-eIF4A1 (Phe163Leu-Ile199Met) eIF4A1SINI cell line. Cells were cultured in tetracycline-free medium. β-Actin was used as the loading control. b, Metagene plot of the 5′ ends of cytoplasmic footprints around start or stop codons in DMSO-treated naïve cells. RPM: reads per million mapped reads. c, Cumulative distribution of changes in the translation of all mRNAs and TOP mRNAs upon treatment with 10 µM PP242 in naïve cells. The p-value was calculated by the two-sided Mann-Whitney U test. d, Western blotting of LARP1, 4EBP1, and phosphorylated 4EBP1 (S65) upon 1 µM and 10 µM PP242 treatment. β-Actin was used as the loading control. e, Western blotting of RPTOR associated with 3 × FLAG-6 × His-tagged LARP1. Cells were subjected to amino acid starvation for 6 h. f, Relative luminescence produced by TOP reporters in naïve, EIF4A1 KO, and SBP-EIF4A2 cells treated with 150 nM Torin-1 for 2 h. The Renilla luciferase signal was normalized to that of firefly luciferase. Data from three replicates (points) and the mean values (bars) are shown. The p-values were calculated by the two-sided Tukey-Kramer multiple comparison test. Tet: tetracycline. g, Western blotting of LARP1, eIF4A1, and 4EBP1 upon the knockdown of LARP1. Cells were treated with 3 µM PP242 for 2 h. β-Actin was used as the loading control. h and i, Relative luminescence produced by the TOP (h) and TOPm (i) reporters in naïve and EIF4A1 KO cells upon the knockdown of LARP1. Cells were treated with 3 µM PP242 for 2 h. The Renilla luciferase signal was normalized to the firefly luciferase signal. Data from three replicates (points) and the mean values (bars) are shown. The p-values were calculated by the two-sided Tukey-Kramer multiple comparison test.

Extended Data Fig. 4 Translational repression upon mTOR inhibition is attenuated in EIF4A1 KO cells, Related to Fig. 4.

a, Global protein synthesis rate in EIF4A1 KO and SBP-EIF4A2 cells, as measured by OP-puro labeling. The quantification of nascent proteins (IRDye 800 signal) normalized to total protein (CBB signal) from three replicates (points) and the mean values (bars) are shown on the right. The p-values were calculated by two-sided Dunnett’s multiple comparison test. Tet: tetracycline. b, Relative viability of naïve, EIF4A1 KO, and SBP-EIF4A2 cells. Tetracycline (Tet) was added 48 h before measurement of luminescence. Data from three replicates (points) and the mean values (bars) are shown. The p-values were calculated by two-sided Tukey-Kramer multiple comparison test. c, Relative luminescence produced by TOP reporters in the indicated cells with the indicated treatment. For the drug treatment, cells were incubated with 0.1 µM hippuristanol for 2 h. The Renilla luciferase signal was normalized to that of firefly luciferase. Data (compared to those in naïve cells with DMSO treatment) from three replicates (points) and the mean values (bars) are shown. d, Relative luminescence produced by TOP reporters in the indicated cells with the indicated treatment. For the drug treatment, cells were incubated with 0.1 µM hippuristanol and/or 3 µM PP242 for 2 h. The Renilla luciferase signal was normalized to that of firefly luciferase. Data (compared to cells without PP242 treatment under each condition) from three replicates (points) and the mean values (bars) are shown. The p-values were calculated by the two-sided Tukey-Kramer multiple comparison test.

Extended Data Fig. 5 Translational repression upon amino acid starvation is attenuated in EIF4A1 KO cells, Related to Fig. 4.

a, Phosphorylation status of mTORC1 substrates (S6K1 and 4EBP1) and an mTORC2 substrate (AKT) in naïve cells upon amino acid starvation for 6 h. Cells were treated with PP242 or Torin-1 for 2 h. β-Actin was used as the loading control. b, Polysome analysis of naïve and EIF4A1 KO cells upon amino acid starvation for 6 h. c, Western blotting of LARP1, RPS6, and RPL11 proteins from subpolysome, light polysome, and heavy polysome fractions. d, Abundances of endogenous TOP mRNAs (RPL10 and RPL13) from subpolysome, light polysome, and heavy polysome fractions, as measured by RT-qPCR. The relative poly/subpoly ratio [(light+heavy)/sub] is shown on the right. The p-values were calculated by two-sided Welch’s t-test. e, Boxplot of changes in the levels of all mRNAs (n = 13,300 mRNAs) and TOP mRNAs (n = 93 mRNAs) upon EIF4A1 knockout. The medians (center line), upper/lower quartiles (box limits), and 1.5× interquartile ranges (whiskers) are shown. The p-values were calculated by the two-sided Mann-Whitney U test.

Extended Data Fig. 6 LARP1 protein purified from PP242-treated cells exhibits increased interaction for the TOP motif, Related to Fig. 5.

a, CBB staining of recombinant 3 × FLAG-6 × His-LARP1 purified from HEK293 cells treated with or without 10 µM PP242. b, (Top) An RNA crosslinking assay was performed to investigate the interactions between recombinant LARP1 and 0.4 µM cap-labeled TOP or TOPm RNAs. 3 × FLAG-6 × His-LARP1 purified from HEK293 cells treated with or without 10 µM PP242 was used. (Bottom) Input cap-labeled TOP or TOPm RNAs were spotted on the nylon membrane. c, SYPRO Ruby staining of recombinant 3 × FLAG-6 × His-LARP1 and LARP1(∆DM15) purified from HEK293 cells treated with 10 µM PP242. d, RNA crosslinking assay of the interaction between purified LARP1(∆DM15) protein and 0.4 µM cap-labeled TOP RNA. 3 × FLAG-6 × His-LARP1 and LARP1(∆DM15) purified from PP242-treated cells were used. e, CBB staining of recombinant 3 × FLAG-6 × His-LARP1 and LARP1(∆327-496) purified from HEK293 cells treated with 10 µM PP242. f, Western blotting of LARP1 associated with SBP-eIF4A1 proteins. Cells were subjected to amino acid starvation for 6 h. g, In vitro pulldown assay of the interactions between recombinant eIF4A1 and LARP1 purified from cells treated with or without 10 µM PP242. Unconjugated beads and beads conjugated to 3 × FLAG-6 × His-tagged LARP1 were incubated with 2 µM or 10 µM His-tagged eIF4A1. For the loading controls, 2 µM and 10 µM 6 × His-eIF4A1 protein were used. Proteins were stained using SYPRO Ruby.

Extended Data Fig. 7 The decrease in cell viability upon mTOR inhibition is attenuated in EIF4A1 KO cells, Related to Fig. 6.

a, Relative viability rate of naïve, EIF4A1 KO, and SBP-EIF4A2 cells treated with Torin-1 for 48 h. Tetracycline (Tet) was added simultaneously with Torin-1. Signals were normalized to those in DMSO-treated samples. Mean values and error bars indicating standard deviations (n = 3 independent replicates) are shown. The p-values for 0.04 µM Torin-1 were calculated by the two-sided Tukey-Kramer multiple comparison test. b, Relative viability rate of naïve, EIF4A1 KO, and SBP-EIF4A2 cells upon amino acid starvation for 30 h. Signals were normalized to those in nonstarved samples. Data from three replicates (points) and the mean values (bars) are shown. The p-values were calculated by the Tukey-Kramer multiple comparison test. Tet: tetracycline. c, Change in the cell numbers of Eif4a1 KO MEFs. A total of 1 × 104 cells were incubated with 1 µM tamoxifen for 24 h and were then cultured in medium without tamoxifen for the indicated time. Mean values and error bars indicating standard deviations (n = 3 independent replicates) are shown. The p-values at 4 days were calculated by two-sided Welch’s t-test.

Extended Data Fig. 8 Characterization of EIF4A2 KO cells, Related to Fig. 7.

a, CBB staining of purified SBP-eIF4A2 protein for RNA pulldown-Seq analysis. b, Electropherogram of RNAs from the lysates used to obtain the data shown in Fig. 7b. c, Schematic of gRNAs designed for EIF4A2 gene knockout and mutation mediated by CRISPR/Cas9 gene editing. d, Western blotting of eIF4A1, eIF4A2, eIF4G1, and eIF4E1 in naïve, EIF4A1 KO, and EIF4A2 KO cells. β-Actin was used as the loading control. e, Global protein synthesis rate in EIF4A1 KO and EIF4A2 KO cells, as measured by OP-puro labeling. The quantification of nascent proteins (IRDye 800 signal) normalized to total protein (CBB signal) in three replicates (points) and the mean values (bars) are shown on the right. The p-values were calculated by two-sided Dunnett’s multiple comparison test. f, Viability of naïve, EIF4A1 KO, and EIF4A2 KO cells. Mean values and error bars indicating standard deviations (n = 3 independent replicates) are shown.

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables 1

Supplementary Tables 1–7 with results of RNA pulldown-seq, proteomics, ribosome profiling, RNA-seq and reporter assays.

Source data

Source Data Fig. 1

Uncropped western blots and gels.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shichino, Y., Yamaguchi, T., Kashiwagi, K. et al. eIF4A1 enhances LARP1-mediated translational repression during mTORC1 inhibition. Nat Struct Mol Biol (2024). https://doi.org/10.1038/s41594-024-01321-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41594-024-01321-7

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing